Next Article in Journal
Ultrasonically Prepared Sodium Heparin-Stabilized Indocyanine Green/Nano-Hydroxyapatite Suspension for Collaborative Photodynamic and Photothermal Tumor Therapy
Next Article in Special Issue
Hybrid Perovskites and 2D Materials in Optoelectronic and Photocatalytic Applications
Previous Article in Journal
Indentation Size Effect in Electrodeposited Nickel with Different Grain Size and Crystal Orientation
Previous Article in Special Issue
Micro-Structure and Dielectric Properties of Ti3C2Tx MXene after Annealing Treatment under Inert Gases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxidized Graphitic-C3N4 with an Extended π-System for Enhanced Photoelectrochemical Property and Behavior

1
Institute for Advanced Materials and Technology, University of Science and Technology Beijing, Beijing 100083, China
2
National Materials Corrosion and Protection Data Center, University of Science and Technology Beijing, Beijing 100083, China
3
BRI Southeast Asia Network for Corrosion and Protection (MOE), Shunde Innovation School, University of Science and Technology Beijing, Foshan 528399, China
4
Institute of Powder Metallurgy and Advanced Ceramics, School of Materials and Engineering, University of Science and Technology Beijing, Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(9), 1386; https://doi.org/10.3390/cryst13091386
Submission received: 24 August 2023 / Revised: 13 September 2023 / Accepted: 16 September 2023 / Published: 18 September 2023

Abstract

:
In this work, an oxidized g-C3N4 film was successfully synthesized using a two-step acid treatment and electrophoretic deposition method. The delocalized π-system of the oxidized g-C3N4 film was extended via an annealing treatment. We investigated the influence of hydrogen bonding reversibility and the oxidation treatment of g-C3N4 on the photoelectrochemical property and photocathodic protection for 304 stainless steel (304 SS). The resulting oxidized g-C3N4 photoelectrode with an extended π-system presents a remarkably enhanced photogenerated electron transfer capability from the photoelectrode to 304 SS (photoinduced OCP negative shift of −0.55 VAgCl) compared with oxidized g-C3N4 and protonated g-C3N4. The oxidation of g-C3N4 facilitates the formation of a porous structure and the introduction of abundant oxygen functional groups, which could promote the effective separation and transport of photogenerated electron–hole pairs. The hydrogen bonding reversibility contributes to the extension of the delocalized π-conjugation system, which could enhance light absorption efficiency. Meanwhile, the annealing treatment is beneficial for prolonging the lifetime of photoelectrons, which could reduce the recombination rate of charge carriers. In addition, to understand how the oxidation treatment and annealing treatment affect the charge transfer behavior, the electronic band structure was investigated, and we found that the oxidized g-C3N4 film with an extended π-system possesses a more negative conduction band position, which could reduce the energy barrier of the photogenerated electron transfer.

1. Introduction

Photocatalysis is regarded as a promising green and sustainable technology for the solution to the issues of the energy crisis and environmental pollution [1,2,3]. Photocathodic protection applied in the corrosion of metals is one of the important research branches in photocatalysis. It can utilize renewable solar energy to realize the protection of metallic materials through the transfer of the photo-generated electrons provided by semiconductor photocatalysts to metals. Because of the solar energy without cost and the photocatalysts without consumption in the protection process, it is an ideal economically and environmentally friendly anticorrosion technology, possessing great potential for future development and application. There has thus been significant effort devoted to the development of highly efficient catalysts for the photoelectrochemical protection of metals [4,5,6].
As a typical metal-free organic semiconductor, graphitic carbon nitride (g-C3N4) has gained wide attention as a potential photocatalyst because of its appropriate electronic structure, suitable band gap (ca. 2.7 eV) for solar light utilization, chemical and thermal stability, and low cost [7,8,9]. Especially in the photocathodic protection of metals, the negative conduction band level of g-C3N4 makes it easier to improve the energy barrier between the semiconductors and the protected metals [10,11,12]. However, the photocathodic protection performance of pristine g-C3N4 still requires improvement, due to the low light absorption and the poor charge transport and separation. In order to address these problems, a number of efforts have been made to modify g-C3N4. Zhang et al. [13] reported the synthesis of g-C3N4 homojunction films and found that constructing a homojunction contributes to enhanced charge separation, promoting the photocathodic protection of 316 stainless steel and Q235 carbon steel. Jing et al. [14] developed a K and I co-doped g-C3N4 photoelectrode for the photoelectrochemical cathodic protection of 316 L stainless steel, proving the important role of K doping modulating the band structure of g-C3N4 and promoting the charge transfer. Ma et al. [11] successfully constructed a Fe2O3/g-C3N4 anti-corrosion coating via conductive adhesive, exhibiting unidirectional electron transport from g-C3N4 to the metal matrix and excellent photocathodic protection performance. Furthermore, in recent years, oxidized g-C3N4, characterized by an enlarged surface area and electron-rich states, has attracted research interest in the photocatalysis field [15,16,17,18]. It has been proved that the oxidation of g-C3N4 could remarkably improve photocatalytic activity. Yang et al. [19] utilized the thermal oxidation method to develop oxidized porous g-C3N4 for enhanced photocatalytic hydrogen production. Dao et al. [20] utilized the chemical oxidation of a K2Cr2O7/H2SO4 mixture to prepare an oxidized g-C3N4 and studied the boosted photocatalytic hydrogen evolution of Pt/oxidized g-C3N4. Speltini et al. [21] obtained oxidized g-C3N4 via the hot HNO3 oxidation method for improved photocatalytic H2 production assisted by aqueous glucose biomass. Despite the merits of oxidized g-C3N4 as photocatalysts in degradation and hydrogen production, few studies have focused on oxidized g-C3N4, especially oxidized g-C3N4 film, for the photocathodic protection of metals. Meanwhile, the underlying influence of hydrogen bonding reversibility and the oxidation treatment of g-C3N4 on the charge transfer of photocathodic protection is unclear.
Inspired by the above discussion, in this work, we report an oxidized g-C3N4 photoelectrode with an extended π-system to boost the photoelectrochemical property and photocathodic protection for 304 SS. The oxidized g-C3N4 film was successfully prepared through HCl protonation, HNO3 oxidation, and electrophoretic deposition methods. The delocalized π-system of the oxidized g-C3N4 was extended through a one-step annealing treatment. The crystalline structures, morphologies, chemical composition, optical properties, and band edge position of the resulting photoelectrodes were investigated. The effects of oxidation treatment and delocalized π-system extension on the photoelectrochemical property and behavior of the g-C3N4 film obtained by electrophoretic deposition were studied. The promoting mechanism of the oxidized g-C3N4 photoelectrode with an extended π-system in the photocathodic protection of 304 SS was demonstrated.

2. Experimental Section

2.1. Material Preparations

2.1.1. Synthesis of Oxidized g-C3N4 Powders

All of the reagents were used without further purification. The deionized water was treated at 100 °C before use.
Firstly, the bulk g-C3N4 powders were synthesized according to previous reports [15]. More specifically, mixed powders of 4 g melamine (AR, Sinopharm, Shanghai, China) and 1 g thiourea (AR, Sinopharm, Ltd., Shanghai, China) were calcined under air at 570 °C in a muffle furnace for 4 h with a heating rate of 2 °C/min, and the bulk g-C3N4 was obtained and designated as B-CN.
Secondly, 0.1 g of B-CN was dispersed into 200 mL of 0.5 mol/L HCl (36~38%, AR, Sinopharm, Shanghai, China) aqueous solution under ultrasonic treatment for 30 min and then continuously stirred for 2 h. The precipitate was collected by filtration, washed with deionized water until the pH of the solution reached 7, and dried at 60 °C for 12 h. The protonated g-C3N4 was obtained and designated as P-CN.
Then, 0.1 g of P-CN samples and 10 mL HNO3 (65~68%, GR, Sinopharm, China) were added to a 50 mL round-bottom flask, stirred at 350 rpm, and refluxed at 80 °C for 2 h. After cooling down to room temperature, the suspension was washed with 0.01 mmol/L NaOH (AR, Sinopharm, China) aqueous solution until the pH of the solution reached 7, followed by washing with deionized water and ethanol (AR, Sinopharm, Shanghai, China) several times and freeze drying for 24 h. Finally, the oxidized g-C3N4 was obtained and designated as OP-CN.

2.1.2. Electrophoretic Deposition of OP-CN Films and Annealing Treatment

Fluorine-doped tin oxide (FTO) conductive glass (14 Ω/cm2, Wuhan Lattice Solar, Wuhan, China) was ultrasonically cleaned in acetone (AR, Sinopharm, Shanghai, China), ethanol, and deionized water for 0.5 h sequentially.
The obtained P-CN and OP-CN powders were electrophoretically deposited on the surface of the FTO glass, and the typical electrophoretic procedure was as follows:
First, 40 mg of P-CN (or OP-CN) powders and 10 mg of iodine powder (AR, Aladdin, Shanghai, China) were dispersed into 50 mL of acetone solution under stirring for 1 h to obtain a uniformly dispersed slurry. Electrophoretic deposition was performed on a two-electrode configuration using the slurry as the electrolyte. The FTO glass (10 × 30 mm) and a platinum plate served as the cathode and anode, respectively, with a distance of 15 mm. The P-CN (or OP-CN) samples were deposited on the conductive surface of FTO at 25 V of applied potential for 5 min. They were further dried in a vacuum at 120 °C for 12 h.
In order to increase the delocalized π-conjugation system of OP-CN, the obtained OP-CN film was annealed at 350 °C for 1 h, and the obtained film on the surface of the FTO substrate was denoted as OP-CN(A). The synthesis conditions for the above three film specimens are listed in Table 1.

2.2. Characterization

The X-ray diffraction (XRD) pattern of samples was obtained using a Rigaku SmartLab X-ray diffractometer (Cu Kα radiation, Rigaku, Tokyo, Japan) over a 2θ range from 10 to 50°. X-ray photoelectron spectroscopy (XPS, Kratos Axis Ultra DLD spectrometer, Al Kα X-ray source, Shimadzu, Kyoto, Japan) was used to determine the surface elemental compositions. The morphology and nanostructure characteristics of the resulting samples were determined using a FEI Tecnai G2 F20 high-resolution transmission electron microscope (TEM, FEI, Oberkochen, Germany) and a ZEISS GeminiSEM 300 field-emission scanning electron microscope (FE-SEM, ZEISS, Oberkochen, Germany). Diffuse UV-Vis absorption (DRS UV-Vis) spectra were obtained using a Shimadzu spectrophotometer (UV-2700 Shimadzu, Kyoto, Japan) with the integral sphere in the range of 300–600 nm. Fourier transform infrared (FT-IR) spectroscopic analysis was carried out using a Thermo Scientific Nicolet iS20 spectrometer (Thermo Fisher Scientific, Waltham, MA, USA).

2.3. Photoelectrochemical Measurements and Photoelectrochemical Cathodic Protection Measurements

The photoelectrochemical tests and photocathodic protection experiments were performed on a CHI660E electrochemical workstation (CHI, Shanghai, China) in a typical three-electrode configuration. A 300 W Xe lamp (CEL-HXF300, Beijing Perfectlight Technology Co., Ltd., Beijing, China) with an AM 1.5 filter (100 mW/cm2) served as the simulated solar light irradiation. For the photoelectrochemical tests, 0.5 mol/L NaOH aqueous solution was used as the electrolyte. The prepared film photoelectrode served as the working electrode. A platinum plate (1 × 1 cm2) and an Ag/AgCl electrode were employed as the counter electrode and reference electrode, respectively. The photocurrent density at the open circuit potential was measured under irradiation by a 300 W Xe lamp. EIS measurements were conducted at the open circuit potential over the frequency range of 0.01 to 105 Hz under light irradiation. Mott–Schottky data were collected at a scan rate of 50 mV/s with a frequency of 1 kHz. In addition, the photocathodic protection performance of the prepared films was investigated by measuring the OCP curves of the protected 304 SS coupled with the photoelectrode, and the corresponding experimental device is shown in Figure S1, in which the electrolyte of the corrosion cell was 3.5 wt% NaCl solution and that of the photoelectrochemical cell was 0.15 M Na2S + 0.1 M Na2SO3 solution.

3. Results and Discussion

3.1. Characterization of the Oxidized Graphitic Carbon Nitride Powders

The synthesis process for the oxidized g-C3N4 (OP-CN) samples is depicted in Scheme 1. First, the bulk g-C3N4 (B-CN) powders were prepared using a previously reported thermal polymerization method. Subsequently, the yellow bulk powders were exfoliated to obtain the protonated g-C3N4 (P-CN) samples using the HCl acidification–sonication method. Finally, the exfoliated P-CN was oxidized using a strong oxidizing acid (HNO3) to obtain the oxidized g-C3N4 (OP-CN) sample. The crystal structure of B-CN, P-CN, and OP-CN powders as revealed by X-ray powder diffraction (XRD) characterization is shown in Figure 1. Two diffraction peaks at ca. 12.9° and 27.6° were observed for B-CN, P-CN, and OP-CN, which can be indexed as the typical signals of (100) and (002) of polymeric carbon nitride (JCPDS, 87-1526) [22,23]. Moreover, the intensity of diffraction peaks for P-CN evidently increases compared with that of B-CN, which indicates that the appropriate protonation of HCl can improve the crystallinity of g-C3N4. Protonation to open the edges in the CN framework through broken structural hydrogen bonding could reduce the stacked structure and facilitate the exposure of the lamellar texture [24,25]. However, after HNO3 pretreatment, the diffraction peaks of OP-CN, especially the extremely low (100) peak, present an obvious decrease. This is attributed to the breakage of the hydrogen bonds in the plane periodicity of the aromatic systems because of the strong oxidative conditions of HNO3 [26]. It is worth noting that the (002) diffraction peak of OP-CN shows a positive shift of 0.2° from 27.6° to 27.8°, implying a decrease in the interplanar distance owing to the introduction of O atoms and the disturbance of the graphic structure [16]. The electronegativity of added O atoms is stronger than that of the N atoms in the framework, which could strengthen the interactions between the g-C3N4 lamellar texture, leading to the shortening of the interplanar distance. In order to further confirm the presence of oxygen in OP-CN, XPS survey spectra of B-CN, P-CN, and OP-CN samples are provided in Figure 1b. The visible peaks of three elements, C, N, and O, can be found in the survey spectrum of OP-CN, while the survey spectra of B-CN and P-CN samples show an extremely low O peak and two high C and N peaks with similar intensity to that of OP-CN. This feature further confirms the addition of oxygen to g-C3N4 after oxidation treatment, which is in accordance with the results of XRD.
The morphology and nanostructure of B-CN, P-CN, and OP-CN powders were observed using an SEM and a TEM, and the obtained images are shown in Figure 2. It can be seen from Figure 2a,b that both the B-CN and P-CN samples display the typical lamellar stacked structure. Meanwhile, there are more obvious curl edges in P-CN samples because of the exfoliation of the stacked lamellar structure caused by protonation. As shown in Figure 2c, the OP-CN samples exhibit interconnected porous networks in addition to the lamellar texture. Upon close observation, it can be observed that the particle boundaries of the B-CN and P-CN samples are abundant (inset of Figure 2a,b), while those of the OP-CN samples become indistinct (inset of Figure 2c). The TEM image in Figure 2d depicts the morphology of OP-CN. A thin layered structure with a wrinkle shape is observed, where several black strips could correspond to the overlap layers.
The DRS UV-Vis spectra of B-CN, P-CN, and OP-CN samples are exhibited in Figure 3. Both the B-CN and P-CN samples present very similar optical absorption in the range of 400~600 nm, and their absorption edges are about 477 nm. This indicates that the HCl moderate protonation of B-CN samples in this work has little effect on the absorption edge, and the electronic band structure could remain largely the same. After oxidation treatment, there is the presence of a semiconductive electronic band structure for the OP-CN samples, but the absorption edge shows an evident blue shift from 477 nm to 446 nm. According to the Tauc plot ((αhv)1/2hv) method (inset of Figure 3a), the band gap energies of three samples can be calculated, in which the Eg values of B-CN and P-CN are both 2.60 eV, while that of OP-CN is 2.78 eV with an increase of 0.18 eV. This increase could be mainly induced by the oxidation of the P-CN frameworks, enabling a partially decreased delocalized π-conjugation system [26,27]. In order to study the effect of HNO3 oxidation on the structure of carbon nitride frameworks, the B-CN, P-CN, and OP-CN samples were characterized by FT-IR spectra, and the results are shown in Figure 3b. It could be found that the FT-IR spectrum of P-CN is similar to that of B-CN, and there are three typical characteristic peaks at 810 cm−1, 1200–1650 cm−1, and 3000–3360 cm−1, corresponding to the breaking mode of triazine units, the stretching mode of CN heterocycles, and the stretching vibration of N-H or the surface hydroxyl groups. For the OP-CN samples, the breaking vibration at 809 cm−1 for triazine units and the skeletal vibrations at 1200–1650 cm−1 for CN heterocycles are still present, which demonstrates the excellent chemical inertness of the melon units [10,23,28]. Significantly, the intensities of the peak at 1085 cm−1 (belonging to the stretching vibration of C–O), the peak at 1635 cm−1 (belonging to the carboxyl group and carbonyl group), and the peak at 3154 cm−1 (belonging to the hydroxyl functional groups) for OP-CN are stronger than those of P-CN [29,30]. These results indicate that the intrinsic structure of pristine g-C3N4 can be preserved but with abundant oxygen functional groups on the edge of CN frameworks after HNO3 etching, which is in agreement with the results of XRD and XPS.

3.2. The Extension of the Delocalized π-Conjugation System for OP-CN Film

It has been discussed and proved that the interfacial hydrogen bonding between the melon units in the frameworks of g-C3N4 is disrupted by the strong oxidation of HNO3, thereby promoting the further depolymerization of g-C3N4 to obtain the smaller particles. These smaller particles would facilitate the formation of a uniform film using the electrophoretic method. Meanwhile, the interconnected porous structure after the HNO3 oxidation treatment would be beneficial for the diffusion and mass transfer in film photoelectrodes. However, according to the results obtained using DRS UV-Vis, the oxidation treatment prompted the blue shift of the light absorption edge for P-CN samples, indicating an enlargement of the band gap and a reduction in the light absorbance range for the film photoelectrode. This result leads to a low solar energy conversion efficiency of the OP-CN photoelectrode and devalues its photochemical application.
It is worth mentioning that hydrogen bonds have a strong reversibility and can be easily manipulated [31]. The extended π-conjugation system could be re-established through an annealing treatment, thereby narrowing the band gap of OP-CN and improving its light absorption [26].
In order to prove the above discussion, the OP-CN(A) film was prepared through annealing at 350 °C for 1 h. The corresponding crystal structure and optical properties were characterized by XRD and DRS UV-Vis, and the OP-CN film was used as a reference. As shown in Figure 4a, after deducting the characteristic peaks of SnO2 (JCPDS No. JCPDS No. 41-1445) belonging to the FTO substrate, the data originating from OP-CN(A) film can be well indexed to the diffraction peaks of g-C3N4. It is widely accepted that the (002) and (100) peaks derive from the interlayer reflection of a graphite-like structure and the in-plane repeated tri-s-triazine units, respectively. The OP-CN powders (Figure 1) and film (Figure 4) both present an evident (002) peak and an extremely low (100) peak, which implies that the in-plane periodicity of the aromatic systems has been partially destroyed because of the disruption of the hydrogen bonds by HNO3 treatment. As expected, the (100) peak of OP-CN(A) can be observed compared with OP-CN, suggesting that the OP-CN structure can be recovered to some extent and the extended π-conjugation system can be rebuilt via annealing treatment. In order to further demonstrate the extension of the aromatic π-conjugation system for OP-CN(A) compared with OP-CN, the DRS UV-Vis spectra of OP-CN and OP-CN(A) are exhibited in Figure 4b. The absorption edge of OP-CN(A) is about 475 nm, but there is an obvious red shift from 446 nm to 475 nm compared with OP-CN. This red shift directly indicates that the delocalized π-conjugation system can be extended after annealing treatment [27,32]. The band gap energy of OP-CN(A) can also be estimated as 2.61 eV using the Tauc plot method (inset of Figure 4b).

3.3. Photoelectrochemical and Photocathodic Protection Performance of OP-CN with Extended π-Conjugation Systems

The photoelectrochemical performance of P-CN and OP-CN photoelectrodes obtained by electrophoretic deposition and OP-CN(A) photoelectrode obtained by the annealing treatment of OP-CN film was investigated using the transient photocurrent response curve (I-t) and electrochemical impedance spectroscopy (EIS). It can be clearly observed in Figure 5a that three kinds of photoelectrodes show an evidently positive response to light irradiation in the variation in photocurrent density with on/off illumination. The P-CN photoelectrode presents the lowest photocurrent density of 4 μm/cm2, while the photocurrent density of OP-CN photoelectrode increases to 13 μm/cm2. The enhancement of photocurrent density can be explained as follows: On the one hand, there is the formation of the porous structure. The interconnected porous structure facilitates the diffusion and mass transfer of electrolytes, which can consume the photogenerated holes [19]. On the other hand, there is the introduction of abundant oxygen functional groups. The abundant oxygen functional groups favor the enhancement of the local electric field between the surface and bulk of OP-CN, which can promote the transfer of photogenerated electron–hole pairs [30,33]. Thus, although the absorption edge of OP-CN has a blue shift (Figure 3a), which indicates that its band gap becomes large, the photocurrent density of the OP-CN photoelectrode is higher than that of P-CN because of the formation of porous structure and abundant oxygen functional groups. Further, with the annealing treatment of the OP-CN film, the OP-CN(A) photoelectrode shows a significantly enhanced photocurrent density (48 μm/cm2), which might be mainly attributed to the fact that the extended delocalized π-system decreases the optical gaps.
In order to study charge transfer behaviors under illumination, EIS measurements were carried out at an open circuit voltage under light irradiation. The Nyquist plots and Bode phase angle plots are given in Figure 5b,c. The inset in the Nyquist plots exhibits the equivalent circuit model with a two-time constant used for the plot fitting, and the fitting data are listed in Table 2. In the equivalent circuit of the Nyquist plots [34,35], Rs represents the series resistance, Rf refers to the film resistance, Rct is the charge transfer resistance, and CPE1 and CPE2 represent the constant phase elements. The constant phase element (CPE) can be expressed by the equation C P E = 1 Y 0 ( j w ) n , where Y0 and n represent the admittance magnitude of the CPE and the exponential term (0 < n ≤ 1), respectively; ω is the angular frequency in rad s−1 (ω = 2πf); and j2 = −1 is an imaginary number [36,37]. The order of charge transfer resistance for the P-CN, OP-CN, and OP-CN(A) photoelectrodes is as follows: OP-CN(A) (3.20 kΩ) < OP-CN (6.48 kΩ) < P-CN (9.14 kΩ). The Rct value of the OP-CN(A) photoelectrode is significantly reduced compared with those of the OP-CN and P-CN photoelectrodes. This indicates that HNO3 oxidation can enhance charge carrier transport in g-C3N4, and the subsequent annealing treatment further decreases the defect centers for charge recombination, which is in agreement with the results of the above photocurrent–time tests.
Additionally, the characteristic peaks of the P-CN, OP-CN, and OP-CN(A) photoelectrodes can be observed in Figure 5c in the Bode phase angle plots of the photoelectrodes under illumination. The lifetime of photogenerated electrons (τ) is associated with the frequency value (fmax) corresponding to the first characteristic peak, which can be estimated using the equation τ = 1 2 π f m a x [35,38]. The calculated τ values of the three photoelectrodes are 201 ms for P-CN, 264 ms for OP-CN, and 348 ms for OP-CN(A). As it is known, the prolonged lifetime of photoelectrons would be beneficial for the efficient separation of photogenerated carriers. The OP-CN(A) photoelectrode shows the longest photoelectron lifetime, indicating that the oxidation and subsequent annealing treatment can reduce the recombination rate of charge carriers, The results of the above EIS and photocurrent response tests can illustrate the OP-CN(A) photoelectrode presents the best photoelectrochemical performance compared with P-CN and OP-CN.
The photocathodic protection performance of P-CN, OP-CN, and OP-CN(A) photoelectrode was evaluated by monitoring the variation in photoinduced open circuit potential (OCP). Figure 5d shows the time-dependent curves of OCP with on/off illumination for three photoelectrodes coupled with the 304 SS. The OPC curve of bare 304 SS is about −0.21 V. Three kinds of photoelectrodes all present a sensitivity to light irradiation, and there is a sharp decrease in the photoinduced OCP from the darkness to the illumination, which can be attributed to the photogenerated electrons migrating from the photoelectrode to the coupled 304 SS, polarizing the metals and inducing a negative shift of their corrosion potential. The photoinduced OPC in the fourth cycle is −0.28 V for P-CN, −0.35 V for OP-CN, and −0.55 V for OP-CN(A). It could be found that OP-CN(A) exhibits a more negative shift of the photoinduced potential compared with OP-CN and P-CN. This might be mainly attributed to the enhanced light absorption efficiency and improved photogenerated charge separation efficiency caused by the extended π-system and the optimized morphology and crystal structure, which is consistent with the above results of the characterization analysis and photoelectrochemical performance. Thus, the OP-CN(A) photoelectrode exhibits the best photocathodic protection performance.

3.4. The Promoting Mechanism of OP-CN Photoelectrode with the Extended π-Conjugation Systems

The investigation of the electronic band structure of P-CN, OP-CN, and OP-CN(A) was performed with the combined analysis of the DRS UV-Vis spectra and Mott–Schottky plots. According to the Mott–Schottky relationship, the interface capacitance (C) as a function of electrode potential (V) under depletion conditions was obtained using the equation 1 C 2 = 2 ( V V f b k T / e 0 ) ε ε 0   A 2 e 0 N d , in which e0 is the electronic charge constant (1.6 × 10−19 C); Vfb represents the flat band potential; ε and ε0 are the relative permittivity constant of g-C3N4 and the vacuum permittivity constant (8.86 × 10−14 F/cm), respectively; Nd and A represent the carrier density and area, respectively; and k and T refer to Boltzmann’s constant and the absolute temperature, respectively [39]. As shown in Figure 6a, the positive slope in the linear region of the Mott–Schottky plots indicates that three photoelectrodes all present typical n-type semiconductor behavior. The flat band potential (Vfb) can be determined through the linear extrapolations with C−2 = 0, which is −0.92 VRHE for P-CN, −0.91 VRHE for OP-CN, and −0.97 VRHE for P-CN. It is known that the Vfb is approximately equal to the conduction band (CB) edge for n-type semiconductors [40]. In addition, the Eg values of P-CN, OP-CN, and OP-CN(A) are 2.60, 2.78, and 2.61 eV, respectively. Thus, the valence band (VB) edge potentials of P-CN, OP-CN, and OP-CN(A) can also be estimated as 1.68, 1.87, and 1.64 VRHE, respectively.
Based on the above discussion, a possible promoting mechanism of OP-CN(A) is proposed. The energy band structure of three photoelectrodes is illustrated in Figure 6b. Under light irradiation, OP-CN(A) can be excited, and the electron–hole pairs are generated on the CB and VB of the semiconductors, respectively. The oxidation treatment and annealing treatment cause a change in the electronic band structure of g-C3N4. The extended π-system can narrow the band gap and improve the light absorption; the negative shift of ECB can reduce the energy barrier for photogenerated electron transfer between the photoelectrode and protected metals. The porous structure, abundant oxygen functional groups, and prolonged lifetime of photoelectrons can boost the separation and transfer efficiency of photogenerated electron–hole pairs. In this case, the OP-CN(A) photoelectrode could provide more effective photogenerated electrons for migration to the surface of 304 SS. Therefore, oxidized g-C3N4 with an extended π-system exhibits the best photocathodic protection performance.

4. Conclusions

In summary, we successfully fabricated an oxidized g-C3N4 film through a two-step acid treatment and electrophoretic deposition method and subsequently extended the delocalized π-system of oxidized g-C3N4 film via an annealing treatment. The resulting samples were characterized by XRD, XPS, SEM, TEM, DRS UV-Vis, and FT-IR. The oxidized g-C3N4 film with an extended π-system presents significantly enhanced photoelectrochemical properties and behaviors, as indicated by the photoinduced potential (−0.55 VAgCl) coupled with 304 SS. Through the investigation of the charge transfer behavior and electronic band structure, the enhancement of photocathodic protection performance can be attributed to the synergistic effect of the narrow band gap, pull-up conduction band position, abundant oxygen functional groups, and prolonged lifetime of photoelectrons. This synergistic effect could extend the light absorption range, promote the separation efficiency of photogenerated electron–hole pairs, and reduce the charge recombination rate, which contributes to the transport of photogenerated electrons onto the protected metals. This work provides a reference for the potential application of photocathodic protection materials as anodes to inhibit the corrosion of steel.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst13091386/s1, Figure S1: The experimental device of photoinduced OCP.

Author Contributions

Conceptualization, Y.C.; methodology, Y.C. and Z.D.; formal analysis, Z.D. and Y.W.; investigation, Y.C., Z.D. and K.S.; data curation, K.S. and Y.W.; writing—original draft preparation, Y.C.; writing—review and editing, X.R.; supervision, Y.C. and X.R.; project administration, X.R.; funding acquisition, Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the fund of Guangdong Basic and Applied Basic Research Foundation (2021A1515110741) and fund of Fundamental Research Funds for the Central Universities (Grant No. FRF-GF-19-031B).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, F.; Weng, B. Photocatalytic hydrogen production: An overview of new advances in structural tuning strategies. J. Mater. Chem. A 2023, 11, 4473–4486. [Google Scholar] [CrossRef]
  2. Samanta, B.; Morales-Garcia, A.; Illas, F.; Goga, N.; Anta, J.A.; Calero, S.; Bieberle-Hutter, A.; Libisch, F.; Munoz-Garcia, A.B.; Pavone, M.; et al. Challenges of modeling nanostructured materials for photocatalytic water splitting. Chem. Soc. Rev. 2022, 51, 3794–3818. [Google Scholar] [CrossRef] [PubMed]
  3. Zhou, P.; Navid, I.A.; Ma, Y.; Xiao, Y.; Wang, P.; Ye, Z.; Zhou, B.; Sun, K.; Mi, Z. Solar-to-hydrogen efficiency of more than 9% in photocatalytic water splitting. Nature 2023, 613, 66–70. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, D.; Liu, Y.; Liu, Y.; Chen, F.; Zhang, C.; Liu, B. A review on recent progress in the development of photoelectrodes for photocathodic protection: Design, properties, and prospects. Mater. Des. 2021, 197, 109235. [Google Scholar] [CrossRef]
  5. Momeni, M.M.; Motalebian, M.; Lee, B.K. Photoelectrochemical performance of titania nanotubes codoped by vanadium and chromium for protection of stainless steel: A promising photoanode for continuous protection in the dark. J. Alloys Compd. 2023, 946, 169429. [Google Scholar] [CrossRef]
  6. Chen, X.; Zhou, G.; Wang, X.; Xu, H.; Wang, C.; Yao, Q.; Chi, J.; Fu, X.; Wang, Y.; Yin, X.; et al. Progress in semiconductor materials for photocathodic protection: Design strategies and applications in marine corrosion protection. Chemosphere 2023, 323, 138194. [Google Scholar] [CrossRef]
  7. Deng, A.; Sun, Y.; Gao, Z.; Yang, S.; Liu, Y.; He, H.; Zhang, J.; Liu, S.; Sun, H.; Wang, S. Internal electric field in carbon nitride-based heterojunctions for photocatalysis. Nano Energy 2023, 108, 108228. [Google Scholar] [CrossRef]
  8. Nagella, S.R.; Vijitha, R.; Naidu, B.R.; Rao, K.S.V.K.; Ha, C.S.; Venkateswarlu, K. Benchmarking recent advances in hydrogen production using g-C3N4-based photocatalysts. Nano Energy 2023, 111, 108402. [Google Scholar] [CrossRef]
  9. Chu, X.; Sathish, C.I.; Yang, J.H.; Guan, X.; Zhang, X.; Qiao, L.; Domen, K.; Wang, S.; Vinu, A.; Yi, J. Strategies for improving the photocatalytic hydrogen evolution reaction of carbon nitride-based catalysts. Small 2023, 2302875. [Google Scholar] [CrossRef]
  10. Ma, X.; Ma, Z.; Zhang, H.; Lu, D.; Duan, J.; Hou, B. Interfacial Schottky junction of Ti3C2TxMXene/g-C3N4 for promoting spatial charge separation in photoelectrochemical cathodic protection of steel. J. Photochem. Photobiol. A-Chem. 2022, 426, 113772. [Google Scholar] [CrossRef]
  11. Ma, Y.; Wang, H.; Sun, L.; Liu, E.; Fei, G.; Fan, J.; Kang, Y.M. Unidirectional electron transport from graphitic-C3N4 for novel remote and long-term photocatalytic anti-corrosion on Q235 carbon steel. Chem. Eng. J. 2022, 429, 132520. [Google Scholar] [CrossRef]
  12. Bu, Y.; Ao, J.P. A review on photoelectrochemical cathodic protection semiconductor thin films for metals. Green Energy Environ. 2017, 2, 331–362. [Google Scholar] [CrossRef]
  13. Zhang, X.; Chen, G.; Li, W.; Wu, D. Graphitic carbon nitride homojunction films for photocathodic protection of 316 stainless steel and Q235 carbon steel. J. Electroanal. Chem. 2020, 857, 113703. [Google Scholar] [CrossRef]
  14. Jing, J.; Chen, Z.; Feng, C.; Sun, M.; Hou, J. Transforming g-C3N4 from amphoteric to n-type semiconductor: The important role of pin type on photoelectrochemical cathodic protection. J. Alloys Compd. 2021, 851, 156820. [Google Scholar] [CrossRef]
  15. Ming, L.; Yue, H.; Xu, L.; Chen, F. Hydrothermal synthesis of oxidized g-C3N4 and its regulation of photocatalytic activity. J. Mater. Chem. A 2014, 2, 19145–19149. [Google Scholar] [CrossRef]
  16. Li, J.; Shen, B.; Hong, Z.; Lin, B.; Gao, B.; Chen, Y. A facile approach to synthesize novel oxygen-doped g-C3N4 with superior visible-light photoreactivity. Chem. Commun. 2012, 48, 12017–12019. [Google Scholar] [CrossRef]
  17. Pisanu, A.; Speltini, A.; Vigani, B.; Ferrari, F.; Mannini, M.; Calisi, N.; Cortigiani, B.; Caneschi, A.; Quadrelli, P.; Profumo, A.; et al. Enhanced hydrogen photogeneration by bulk g-C3N4 through a simple and efficient oxidation route. Dalton Trans. 2018, 47, 6772–6778. [Google Scholar] [CrossRef]
  18. Xu, J.; Gao, Q.; Bai, X.; Wang, Z.; Zhu, Y. Enhanced visible-light-induced photocatalytic degradation and disinfection activities of oxidized porous g-C3N4 by loading Ag nanoparticles. Catal. Today 2019, 332, 227–235. [Google Scholar] [CrossRef]
  19. Yang, L.; Huang, J.; Shi, L.; Cao, L.; Yu, Q.; Jie, Y.; Fei, J.; Ouyang, H.; Ye, J. A surface modification resultant thermally oxidized porous g-C3N4 with enhanced photocatalytic hydrogen production. Appl. Catal. B Environ. 2017, 204, 335–345. [Google Scholar] [CrossRef]
  20. Dao, Q.D.; Nguyen, T.K.A.; Dang, T.T.; Kang, S.G.; Nguyen-Phu, H.; Do, L.T.; Van, V.K.H.; Chung, K.H.; Chung, J.S.; Shin, E.W. Anchoring highly distributed Pt species over oxidized graphitic carbon nitride for photocatalytic hydrogen evolution: The effect of reducing agents. Appl. Surf. Sci. 2023, 609, 155305. [Google Scholar]
  21. Speltini, A.; Scalabrini, A.; Maraschi, F.; Sturini, M.; Pisanu, A.; Malavasi, L.; Profumo, A. Improved photocatalytic H2 production assisted by aqueous glucose biomass by oxidized g-C3N4. Int. J. Hydrogen Energy 2018, 43, 14925–14933. [Google Scholar] [CrossRef]
  22. Li, Y.P.; He, J.Y.; Wang, X.J.; Zhao, J.; Liu, R.H.; Liu, Y.; Li, F.T. Introduction of crystalline hexagonal-C3N4 into g-C3N4 with enhanced charge separation efficiency. Appl. Surf. Sci. 2021, 559, 149876. [Google Scholar] [CrossRef]
  23. Liu, Q.; Guo, Y.; Chen, Z.; Zhang, Z.; Fang, X. Constructing a novel ternary Fe(III)/graphene/g-C3N4 composite photocatalyst with enhanced visible-light driven photocatalytic activity via interfacial charge transfer effect. Appl. Catal. B-Environ. 2016, 183, 231–241. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Thomas, A.; Antonietti, M.; Wang, X. Activation of carbon nitride solids by protonation: Morphology changes, enhanced ionic conductivity, and photoconduction experiments. J. Am. Chem. Soc. 2009, 131, 50–51. [Google Scholar] [CrossRef] [PubMed]
  25. Wei, Y.; Wang, Z.; Su, J.; Guo, L. Metal-Free Flexible Protonated g-C3N4/carbon gots photoanode for photoelectrochemical water splitting. Chemelectrochem 2018, 5, 2734–2737. [Google Scholar] [CrossRef]
  26. Zhang, J.; Zhang, M.; Lin, L.; Wang, X. Sol Processing of Conjugated Carbon Nitride Powders for Thin-Film Fabrication. Angew. Chem. Int. Ed. 2015, 54, 6297–6301. [Google Scholar] [CrossRef]
  27. Shi, A.; Li, H.; Yin, S.; Liu, B.; Zhang, J.; Wang, Y. Effect of conjugation degree and delocalized π-system on the photocatalytic activity of single layer g-C3N4. Appl. Catal. B Environ. 2017, 218, 137–146. [Google Scholar] [CrossRef]
  28. Cui, M.; Cui, K.; Liu, X.; Shi, J.; Chen, X.; Chen, Y. Synergistic effect of mesoporous graphitic carbon nitride and peroxydisulfate in visible light-induced degradation of atenolol: A combined experimental and theoretical study. Chem. Eng. J. 2021, 412, 127979. [Google Scholar] [CrossRef]
  29. Sun, X.; Lei, J.; Jin, Y.; Li, B. Long-Lasting and intense chemiluminescence of luminol triggered by oxidized g-C3N4 nanosheets. Anal. Chem. 2020, 92, 11860–11868. [Google Scholar] [CrossRef]
  30. Li, Z.; Meng, X.; Zhang, Z. Fabrication of surface hydroxyl modified g-C3N4 with enhanced photocatalytic oxidation activity. Catal. Sci. Technol. 2019, 9, 3979–3993. [Google Scholar] [CrossRef]
  31. Zhou, Z.; Zhang, Y.; Shen, Y.; Liu, S.; Zhang, Y. Molecular engineering of polymeric carbon nitride: Advancing applications from photocatalysis to biosensing and more. Chem. Soc. Rev. 2018, 47, 2298–2321. [Google Scholar] [CrossRef] [PubMed]
  32. Cao, S.; Yu, J. g-C3N4-Based photocatalysts for hydrogen generation. J. Phys. Chem. Lett. 2014, 5, 2101–2107. [Google Scholar] [CrossRef] [PubMed]
  33. Chang, Y.; Xuan, Y.; Quan, H.; Zhang, H.; Liu, S.; Li, Z.; Yu, K.; Cao, J. Hydrogen treated Au/3DOM-TiO2 with promoted photocatalytic efficiency for hydrogen evolution from water splitting. Chem. Eng. J. 2020, 382, 122869. [Google Scholar] [CrossRef]
  34. Zhou, Z.; Wang, Y.; Li, L.; Yang, L.; Niu, Y.; Yu, Y.; Guo, Y.; Wu, S. Constructing a full-space internal electric field in a hematite photoanode to facilitate photogenerated-carrier separation and transfer. J. Mater. Chem. A 2022, 10, 8546–8555. [Google Scholar] [CrossRef]
  35. Pan, G.; Li, J.; Zhang, G.; Zhan, Y.; Liu, Y. Binder-integrated Bi/BiOI/TiO2 as an anti-chloride corrosion coating for enhanced photocathodic protection of 304 stainless steel in simulated seawater. J. Alloys Compd. 2023, 938, 168469. [Google Scholar] [CrossRef]
  36. Li, H.; Wang, X.; Liu, Y.; Hou, B. Ag and SnO2 co-sensitized TiO2 photoanodes for protection of 304SS under visible light. Corros. Sci. 2014, 82, 145–153. [Google Scholar] [CrossRef]
  37. Motalebian, M.; Momeni, M.M.; Lee, B.K. Novel photoanodes based on (Mo-Cr) co-doped titania nanotube for highly efficient photocathodic protection performance of stainless steel. Appl. Surf. Sci. 2023, 610, 155091. [Google Scholar] [CrossRef]
  38. Guan, Z.C.; Hu, J.; Wang, H.-H.; Shi, H.Y.; Wang, H.P.; Wang, X.; Jin, P.; Song, G.L.; Du, R.G. Decoration of rutile TiO2 nanorod film with g-C3N4/SrTiO3 for efficient photoelectrochemical cathodic protection. J. Photochem. Photobiol. A Chem. 2023, 443, 114825. [Google Scholar] [CrossRef]
  39. Li, X.; Wan, J.; Ma, Y.; Zhao, J.R.; Wang, Y. Mesopores octahedron GCNOX/Cu2O@C inhibited photo-corrosion as an efficient visible-light catalyst derived from oxidized g-C3N4/HKUST-1 composite structure. Appl. Surf. Sci. 2020, 510, 145459. [Google Scholar] [CrossRef]
  40. Jiang, X.; Sun, M.; Chen, Z.; Jing, J.; Feng, C. High-efficiency photoelectrochemical cathodic protection performance of the TiO2/AgInSe2/In2Se3 multijunction nanosheet array. Corros. Sci. 2020, 176, 108901. [Google Scholar] [CrossRef]
Scheme 1. The synthesis process for oxidized g-C3N4 (OP-CN) samples.
Scheme 1. The synthesis process for oxidized g-C3N4 (OP-CN) samples.
Crystals 13 01386 sch001
Figure 1. (a) XRD patterns and (b) XPS survey spectra of B-CN, P-CN, and OP-CN powders.
Figure 1. (a) XRD patterns and (b) XPS survey spectra of B-CN, P-CN, and OP-CN powders.
Crystals 13 01386 g001
Figure 2. SEM images of (a) B-CN, (b) P-CN, and (c) OP-CN; (d) TEM image of OP-CN.
Figure 2. SEM images of (a) B-CN, (b) P-CN, and (c) OP-CN; (d) TEM image of OP-CN.
Crystals 13 01386 g002
Figure 3. (a) DRS UV-Vis spectra and (b) FT-IR spectra of B-CN, P-CN, and OP-CN powders.
Figure 3. (a) DRS UV-Vis spectra and (b) FT-IR spectra of B-CN, P-CN, and OP-CN powders.
Crystals 13 01386 g003
Figure 4. (a) XRD patterns and (b) DRS UV-Vis spectra of OP-CN (before the annealing treatment) and OP-CN(A) (after the annealing treatment) film.
Figure 4. (a) XRD patterns and (b) DRS UV-Vis spectra of OP-CN (before the annealing treatment) and OP-CN(A) (after the annealing treatment) film.
Crystals 13 01386 g004
Figure 5. (a) Photocurrent density–time curves, (b) Nyquist plots, and (c) Bode phase angle plots of the prepared photoelectrodes including P-CN, OP-CN, and OP-CN(A) under light irradiation; (d) photoinduced OCP–time curves of P-CN, OP-CN, and OP-CN(A) photoelectrodes coupled with 304 SS under light irradiation, and the OPC curve of bare 304 SS.
Figure 5. (a) Photocurrent density–time curves, (b) Nyquist plots, and (c) Bode phase angle plots of the prepared photoelectrodes including P-CN, OP-CN, and OP-CN(A) under light irradiation; (d) photoinduced OCP–time curves of P-CN, OP-CN, and OP-CN(A) photoelectrodes coupled with 304 SS under light irradiation, and the OPC curve of bare 304 SS.
Crystals 13 01386 g005
Figure 6. (a) Mott–Schottky plots and (b) the corresponding electronic band structures of P-CN, OP-CN, and OP-CN(A).
Figure 6. (a) Mott–Schottky plots and (b) the corresponding electronic band structures of P-CN, OP-CN, and OP-CN(A).
Crystals 13 01386 g006
Table 1. The synthesis conditions for three film specimens.
Table 1. The synthesis conditions for three film specimens.
Film SpecimenDeposited Colloidal ParticlesFilm-Forming MethodPost-Processing
P-CN P-CN powdersElectrophoretic deposition/
OP-CN O-CN powdersElectrophoretic deposition/
OP-CN(A) O-CN powdersElectrophoretic depositionAnnealing treatment
Table 2. The equivalent circuit model EIS results for different samples.
Table 2. The equivalent circuit model EIS results for different samples.
SampleRs (Ω)CPE1, Y0
(S·secn)
n1
(0 < n1 ≤ 1)
Rf (Ω)CPE2, Y0
(S·secn)
n2
(0 < n2 ≤ 1)
Rct (kΩ)
P-CN2.473.83 × 10−40.84727.63.30 × 10−40.9939.14
OP-CN1.675.91 × 10−40.849131.914.76 × 10−416.48
OP-CN(A)1.3110.3 × 10−40.82527.18.73 × 10−413.20
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chang, Y.; Dai, Z.; Suo, K.; Wang, Y.; Ren, X. Oxidized Graphitic-C3N4 with an Extended π-System for Enhanced Photoelectrochemical Property and Behavior. Crystals 2023, 13, 1386. https://doi.org/10.3390/cryst13091386

AMA Style

Chang Y, Dai Z, Suo K, Wang Y, Ren X. Oxidized Graphitic-C3N4 with an Extended π-System for Enhanced Photoelectrochemical Property and Behavior. Crystals. 2023; 13(9):1386. https://doi.org/10.3390/cryst13091386

Chicago/Turabian Style

Chang, Yue, Zhongkui Dai, Kaili Suo, Yuhang Wang, and Xiaona Ren. 2023. "Oxidized Graphitic-C3N4 with an Extended π-System for Enhanced Photoelectrochemical Property and Behavior" Crystals 13, no. 9: 1386. https://doi.org/10.3390/cryst13091386

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop